Sei sulla pagina 1di 7

Journal of Membrane Science 392–393 (2012) 88–94

Contents lists available at SciVerse ScienceDirect

Journal of Membrane Science


journal homepage: www.elsevier.com/locate/memsci

Transport of liquid water through Nafion membranes


Qiongjuan Duan a , Huaping Wang a , Jay Benziger b,∗
a
College of Material Science and Engineering, State Key Laboratory for Modification of Chemical Fibers and Polymer Materials, Donghua University, Shanghai 201620, PR China
b
Chemical and Biological Engineering, Princeton University, Princeton, NJ 08544 USA

a r t i c l e i n f o a b s t r a c t

Article history: The flux of liquid water through Nafion membranes of different thickness and equivalent weight was mea-
Received 3 October 2011 sured as a function of hydrostatic pressure and temperature. Hydraulic water transport across Nafion
Received in revised form 4 December 2011 membranes increases with temperature and equivalent weight of the Nafion. Hydraulic permeability
Accepted 5 December 2011
increases with temperature due to both decreased water viscosity and increased hydrophilic volume
Available online 13 December 2011
fraction. Convective flow from the applied hydrostatic water pressure is an order of magnitude greater
than the estimated diffusive water flux associated with the water activity gradient. Water sorption and
Keywords:
hydraulic permeability data predict a hydrophilic pore network with hydrophilic domains 2.5 nm in diam-
Hydraulic permeation
Nafion
eter spaced 5.5 nm apart. The pore network structure from water sorption and hydraulic permeability
Pore network are consistent with the spacing between hydrophilic domains observed with small angle X-ray scattering
experiments.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction by a hydrophobic matrix of tetrafluoroethylene and perfluoro


ethers. Water sorption swells the hydrophilic domains providing
Water transport through Nafion membranes is essential for the paths for proton transport and water diffusion [19,20]. We expect
successful operation of Polymer Electrolyte Membrane (PEM) Fuel that convective water transport driven by a hydraulic pressure will
Cells. Water made at the cathode catalyst layer, situated between also be through the hydrophilic domains. Several versions of Nafion
the PEM and the porous cathode gas diffusion layer (GDL) must membranes are available that differ by membrane thickness, den-
be removed. Depending on the resistances to transport, water can sity of sulfonic acid groups and the capping groups on the ends
either flow through the porous GDL to the cathode gas flow channel of the polymer chains. We anticipate that the hydraulic perme-
or flow through the PEM to the anode catalyst layer, anode GDL and ation of water will correlate with the hydrophilic volume fraction
anode gas flow channel. Water is transported through the PEM both and inversely with the membrane thickness. Recent studies charac-
by diffusion, due to an activity gradient, and by convection, due to terizing the diffusive transport of water across Nafion membranes
a hydraulic pressure difference between the cathode and anode. with different equivalent weights (EW = mass of polymer/mole of
In a typical PEM fuel cell there might be a pressure difference of SO3 − X+ ) and different membrane thickness have shown that inter-
0.1 MPa between the cathode and the anode. There has been many facial transport at the vapor/membrane interface can limit water
studies of water transport by diffusion through PEMs over the past transport across thin Nafion membranes, but there is negligible
several years [1–16]. In contrast there has been only a few reports of transport resistance at the liquid/membrane interface [1,2,4,20,21].
convective water transport through PEMs [2,5,17]. We report here PEM fuel cells are often operated with pressure differences
results for water uptake and convective water transport through between the anode and cathode; water can be transported across
the PEM Nafion. the membrane by diffusion driven by a water activity difference and
Nafion is the most common and popular PEM employed in fuel by convection driven by a hydraulic pressure difference. Two recent
cells. It is a perfluorosulfonic acid made by copolymerization of papers have reported hydraulic permeability of water through
tetrafluoroethylene and a perfluorovinyl ether terminated with a Nafion was small and increased with temperature [2,4]. Kientiz
sulfonyl fluoride [18]. After polymerization the sulfonyl fluoride et al. examined Nafion 212 and 211 and suggested that hydraulic
is hydrolyzed and ion exchanged to produce the perfluorosulfonic permeability was small relative to diffusive transport of water [2].
acid. Nafion microphase separates into hydrophilic domains com- We have extended the previous studies to compare water perme-
prised of sulfonic acid groups and sorbed water that is surrounded ation through Nafion membranes with different thickness, different
equivalent weight and different capping groups. The results show
that hydraulic permeability increases with temperature and vol-
∗ Corresponding author. Tel.: +1 609 258 5416. ume fraction of water in the membrane. The hydraulic radius
E-mail address: benziger@princeton.edu (J. Benziger). of the water channels and their spacing were determined from

0376-7388/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2011.12.004
Q. Duan et al. / Journal of Membrane Science 392–393 (2012) 88–94 89

Fig. 1. Schematic of flow cell for water convection through Nafion membranes.

measurements of water volume fraction and water flow rate in a


Fig. 2. Water flux through Nafion 115 membranes at different temperatures.
porous network. The distance between hydrophilic domains was
∼5.5 nm in agreement with the characteristic length observed by
scattering experiments [18]. membranes were placed in a sealed sample cell constructed from
two copper plates with mica windows and a Viton gasket between
2. Experimental them. Membrane water contents, , were determined gravimet-
rically on samples having the same treatment as those being
2.1. Flow experiments mounted in the SAXS cell.
The sample cell was placed in the path of the X-ray beam. In
Liquid water flow rates were measured with the permeation order to minimize membrane water loss during the data collec-
cell shown schematically in Fig. 1. Two polycarbonate blocks tion, helium gas was used in the flight tube rather than a vacuum.
were machined with plenums 2.5 cm diameter × 1.25 cm deep. The The 1.54 Å Cu-K␣ X-rays were generated by a Philips XRG-3000
blocks were bolted together with a polymer membrane placed sealed tube generator source. The beam was slit collimated and the
between the two plenums. A support cross was fitted in the low scattering was detected by an Anton-Paar compact Kratky cam-
pressure plenum to limit deflection of the polymer membrane. era equipped with a Braun OED-50M detector. Samples were run
Water was fed to one plenum from a pressurized reservoir. The at room temperature for 10 min. Empty beam scattering, sample
hydrostatic pressure from the height of the water reservoir above transmittance, and detector response were corrected for in the
the membrane and the applied gas pressure were combined to data analysis. The data reduction and desmearing procedures are
give a total hydraulic pressure. The outlet from the second plenum described in detail by Register [22]. The invariant scattering inten-
went into a 1.6 mm ID tygon tube. The water flow rate across the sity (q2 I) is plotted as a function of scattering angle or distance
membrane was determined from the distance the water moved and the Bragg spacing is determined by the location of the peak in
through the tygon tube as a function of time. Water flow rates scattering intensity.
were measured at a fixed hydraulic pressure for a period of 2–3 h to
obtain steady state flow rates. Flow rates were measured for both 2.4. Water uptake
increasing and decreasing pressure changes. The permeation cell
was placed inside an insulated heated box where the temperature The equilibrium water uptake was measured as a function of
was controlled. water pressure at different temperatures using an isometric system
[19,23]. A PEM is placed in a fixed volume container. The PEM sam-
2.2. Membrane preparation and treatment ple was evacuated to below 1 Pa (<0.01 mbar) at 80 ◦ C to remove
all the water from the membrane. The pressure container was then
All membranes were treated by boiling in 3% H2 O2 for 1 h, rins- cooled to the desired temperature under vacuum. The valve to the
ing in boiling deionized water for 20 min, boiling in 1 M H2 SO4 vacuum line was shut off and aliquots of water are injected into
for an hour and finally rinsing again in boiling deionized water the pressure vessel and allowed to equilibrate with the Nafion. The
for 30 min. Commercial extruded Nafion 115, Nafion 1110, Nafion resulting pressure is equal to the partial pressure of water. The dif-
1035 and Nafion 212 were obtained from Ion Power (New Castle ference between the quantity of water injected and the water in the
DE) and treated as outlined above. Nafion 115 and Nafion 1110 are gas phase is equal to the amount of water absorbed by the Nafion.
both equivalent weight 1100 with dry thicknesses 127 and 254 ␮m, Saturation was evident when the injection of a liquid aliquot did
respectively. Nafion 212 is equivalent weight 1100 with a dry thick- not cause any change in the pressure.
ness of 51 ␮m. Nafion 212 has a different capping group on the
polymer chain than Nafion 115 and Nafion 1110. Nafion 1035 is 3. Results
equivalent weight 1000 with a dry membrane thickness of 89 ␮m.
The water fluxes as functions of applied hydraulic pressure for
2.3. SAXS sample preparation Nafion 115 at room temperature (∼23 ◦ C), 40, 60 and 80 ◦ C are
shown in Fig. 2. Data for different Nafion membranes, N115, N1110,
Small angle X-ray scattering was carried out on samples of N1035 and N212 at room temperature are shown in Fig. 3. These
Nafion 1110 equilibrated with water vapor and liquid water at data are all well represented by a linear increase of water flux with
room temperature and a sample placed in boiling water. The applied hydraulic pressure. Table 1 summarizes the slope of the
90 Q. Duan et al. / Journal of Membrane Science 392–393 (2012) 88–94

Table 1
Nafion membrane hydraulic permeabilities and water sorption.

Membrane (EW) T (◦ C)  (Pa s) tm (␮m) kw (cm2 × 1016 ) sat = #H2 O/SO3

N115 (1100) 23 .100 127 3.75 21 ± 2


N115 (1100) 40 .065 127 5.04 22 ± 2
N115 (1100) 60 .047 127 5.49 24 ± 2
N115 (1100) 80 .035 127 7.13 25 ± 2
N1110 (1100) 23 .100 254 3.90 21 ± 2
N1035 (1000) 23 .100 89 5.77 21 ± 2
N212 (1100) 23 .100 51 4.09 21 ± 2

water flux (Qw ) vs. pressure (P) for the different samples and dif- increase in water sorption between saturated vapor and liquid, and
ferent temperatures. Membrane permeabilities (kw ), defined by Eq. an additional small increase between saturated liquid at 20 ◦ C and
(1), were calculated from the flux data and are given in Table 1. The 100 ◦ C. The characteristic scattering length increases slightly with
membrane permeability is expected to be independent of water increased water content in the membrane.
viscosity, , and membrane thickness, tm .

kw 4. Discussion
Qw = P (1)
tm
The mechanisms of water transport in Nafion and structure
The results indicate that the three membranes with the 1100 of the “water channels” have been a source of intense interest
equivalent weight, N115, N1110 and N212, all had similar perme- over the past 25 years. Modeling efforts have examined different
abilities circa 4 × 10−16 cm2 at 23 ◦ C. N1035, with an equivalent contributions of water sorption and transport including diffusion
weight of 1000, had permeability about 40% greater than the and convection. Eikerling and co-workers have considered various
1100 EW Nafion. The permeability of N115 membranes increased driving forces for convection including external pressure gradient,
with temperature. Since kw corrects for the viscosity change with capillary pressure, osmotic pressure, and elastic forces associated
temperature, the data show that the permeability increased with with membrane deformation [28–31]. Experiments have lagged
temperature. behind the modeling efforts for water transport and only recently
The saturation uptake of water into the Nafion membranes have experimental results been published that provide details
(reported as  = number of water molecules sorbed per sulfonate of transport resistances [1,4,5,10,20,32,33], mechanical properties
group) are shown in Table 1. The saturation water volume increased [34,35] and water profiles [36–39] as functions of water activity and
modestly with increasing temperature. The saturation water vol- temperature. These recent experimental results can be compared
ume at 23 ◦ C was also larger for the 1000 EW Nafion (N1035) than with previous modeling efforts to improve the understanding of
the 1100 EW Nafion. (Even though  was the same for N115 and structure–property relationships in Nafion.
N1035 the water volume fraction is larger for N1035 because the Water permeation, water sorption and SAXS studies of water in
density of sulfonic acid groups is greater for N1035.) Nafion have been reported in the literature [18]. It is well estab-
Small angle X-ray scattering results for four different Nafion lished that water sorption in Nafion is accompanied by a strong
1110 samples are shown in Fig. 4. More detailed studies can be scattering peak with a characteristic distance of ∼5 nm. Water dif-
found in a number of sources [18,24–27]. The purpose of Fig. 4 is fusion increases with water content in Nafion. However, there
to highlight the characteristic scattering length of ∼5 nm for Nafion has been minimal quantitative connections between the struc-
samples with absorbed water. There is no evidence of any scattering ture of Nafion determined by scattering experiments and transport
in the nm range for proton exchanged Nafion membranes with- rates. We explore here the consistency of convective water trans-
out absorbed water. The other three samples correspond to water port through Nafion and the structure of the hydrophilic domains
absorption at saturation conditions: saturated vapor at room tem- determined from scattering experiments. It will be shown that con-
perature, saturated liquid at room temperature and saturated liquid vective water transport coupled with water sorption predicts water
at 100 ◦ C. The SAXS results for 1100 EW Nafion indicate a small

Fig. 4. Small angle X-ray scattering of Nafion 1110 membranes in saturated vapor,
Fig. 3. Water flux through different Nafion Membranes at 23 ◦ C. room temperature saturated liquid and saturated liquid at 100 ◦ C.
Q. Duan et al. / Journal of Membrane Science 392–393 (2012) 88–94 91

is convected through a hydrophilic “channel” network consisting of a substantial hydrostatic pressure difference (∼100 kPa) between
2.5 nm diameter water channels spaced 5.5 nm apart; this structure the anode and cathode.
shows quantitative agreement with the results from SAXS. Benziger and co-workers, and Monroe et al. concluded that
interfacial water transport at the membrane vapor interface is
4.1. Effect of temperature and equivalent weight on hydraulic the principal resistance to water transport in pervaporation sys-
permeability tems [1,8,20]. Because the limiting transport resistance is at the
vapor/membrane interface, the water activity in the membrane is
Two recent papers have reported measurements of pressure nearly constant across the membrane, approximately equal to the
driven water convection through Nafion. Weber and collaborators liquid water activity, as shown by Eq. (3), where km is a function of
from Toyota reported permeability coefficients for Nafion 211 and temperature.
Nafion 212 [2]. Their values for 1100 EW Nafion is in the range vapor liquid vapor
(Qpervaporation ) = km (amembrane − aw ) ≈ km (aw − aw ) (3)
of 1–8 × 10−16 cm2 very close to those reported here at 60 and w

80 ◦ C. The permeability they measured at 25 ◦ C was 4 times smaller Saturated liquid activity is unity (aw
saturated liquid
= 1). The activity
than the value we report. Holdcroft and co-workers also measured correction for a pressurized liquid is given by the Poynting correc-
hydraulic water permeability through Nafion 212 [4,5,21]. They tion shown in Eq. (4) (assuming that liquid water is incompressible
also obtained values circa 10−16 cm2 similar to those reported here. over the pressures of interest).
These recent values are almost 100 times less than earlier values    
reported by Bernaradi and Verbrugge [17]. We suspect there were P sat )V̄
sat V̄w (P − Pw w
some difficulties with those early experiments. aw = aw (Pw ) + exp dP = 1 + exp
sat RT RT
Hydraulic permeabilities at 23 ◦ C were almost the same for Pw

N115, N1110 and N212, all of which have an equivalent weight sat )V̄
(P − Pw w
of 1100. These three membranes also sorbed the same number of ≈1+ (4)
RT
water molecules per sulfonic acid groups ( = 21H2 O/SO3 at 23 ◦ C).
The hydraulic permeability of N1035, with EW 1000, is greater than The activity correction to pressurized water at 100 ◦ C from satu-
the permeability of the 1100 EW Nafion membranes. ration (0.1 MPa) to 1 MPa pressure (10 bar) is ∼0.001. Hence, we
Our results, as well as those of Kientiz et al. and Adachi et al., conclude that at low to moderate differential pressures no mea-
found that the hydraulic permeability increased with tempera- sureable change to the water pervaporation flux is expected. This
ture. In all three studies the water flux increased more than that is consistent with pressurized pervaporation experiments by Kien-
predicted by the decrease of water viscosity with temperature, indi- tiz et al. and Adachi et al. (We did similar experiments. Initially
cating that the hydraulic permeability increased with temperature. we observed a small change in the steady state flux for pressur-
Results in Table 1 also show that the number of water molecules ized pervaporation through Nafion. However, when we followed
sorbed per sulfonic acid group increased with temperature, indi- the flux dynamically we discovered the change in flux was due to
cating that the volume fraction of water in Nafion increased with bowing of the membrane due to the pressure differential, which
increasing temperature. The increase in hydraulic permeability increased the interfacial area.)
with temperature and equivalent weight is greater than expected There is no convective water flux across the membrane/vapor
from the increased volume fraction of water. The water volume interface. Since interfacial transport at the vapor/membrane inter-
fraction, ϕw , given in Eq. (2) changes by almost 10% while the face is the limiting resistance to pervaporation, any contribution
hydraulic permeability increased by 50–80%. This difference sug- to water transport by hydraulic permeation in the membrane will
gests there may be structural changes to the hydrophilic domains increase the water concentration at the membrane/vapor inter-
with increased water sorption beyond simple volume expansion. face and reduce the already small water activity gradient across
the membrane.
(Nafion /EW)Vw When liquid water is at both membrane surfaces the interfa-
ϕw = (2)
1 + (Nafion /EW)Vw cial transport resistances are negligible. Unless there is a pressure
(or temperature) gradient across the membrane there is no driv-
4.2. Hydraulic permeation vs. pervaporation ing force for water transport. A differential pressure across the
membrane will drive water transport by diffusion and hydraulic
The recent paper by Kientiz et al. compared the water fluxes permeation. Diffusive permeation is driven by the water activity
from liquid/vapor pervaporation and liquid/liquid hydraulic per- gradient across the membrane. The activity difference for pressur-
meation. They observed that the liquid/vapor pervaporation fluxes ized liquid water on both sides of the membrane is given by the
were significantly greater than the liquid/liquid hydraulic perme- Poynting correction shown in Eq. (5).
ation fluxes. Kientiz et al. suggested that hydraulic permeation  Pcathode     
could be neglected relative to the pervaporation if there was an V̄w (Panode − Pcathode )V̄w
aw = exp dP = exp
activity gradient [2]. Adachi et al. also made similar observations Panode
RT RT
[4]. In our liquid/vapor permeation experiments in our labs we
also observed larger pervaporation water fluxes than the convec- (Panode − Pcathode )V̄w
≈ (5)
tive water fluxes we report here [1,20,32]. Is hydraulic permeation RT
important to water transport in Nafion? The diffusive water flux (Qw,diffusive ) associated with the hydro-
Water transport across Nafion membranes is essential to both static pressure difference across the membrane of thickness tm , is
membrane humidifiers and PEM fuel cells. The water activity dif- given by Eq. (6); where Dw is the diffusion coefficient of water and
ferences across the membrane are quite different in fuel cells cw is the concentration of water in the membrane (cw = NSO3 =
compared to membrane humidifiers. Nafion membrane humid- Nafion /EW where NSO3 is the density of sulfonic acid groups in
ifiers operate as pervaporation cells with a small hydrostatic Nafion). Diffusion coefficients for 1100 EW Nafion were reported
pressure difference (<10 kPa) but large relative humidity difference by Zhao et al. [20]; Dw (aw = 1) ≈ 5 × 10−6 cm2 s−1 .
across the membrane [40–42]. In contrast, fuel cells have saturated
liquid/gas streams at both sides of the membrane throughout most aw Dw NSO3 (Panode − Pcathode )V̄w
Qw,diffusive ≈ Dw cw,mem = (6)
of the fuel cell; there is a negligible relative humidity difference, but tm tm RT
92 Q. Duan et al. / Journal of Membrane Science 392–393 (2012) 88–94

The ratio of the diffusive flux (Eq. (6)) to the convective flux (Eq.
(1)) with liquid present on both sides of the membrane is given by
Eq. (7).

Qdiffusive Dw NSO3 V̄w


= ≈ 3 × 10−2 (7)
Qconvective RTkw
When water is present at both sides of the Nafion membrane the
diffusive flux from an applied pressure is only a small fraction of the
hydraulic flux (∼3%). In PEM fuel cells when the anode and cathode
gas streams are both fully humidified the transport of water across
Nafion membranes will be dominated by hydraulic permeation.

4.3. Water removal from the catalyst layer: through the


membrane or through the GDL?

Water is produced at the cathode catalyst layer. The vapor is


saturated in the cathode flow channel. There is no water activ-
ity gradient to drive diffusion from the cathode catalyst layer to
Fig. 5. Cell model for hydrophilic network structure. (A) This is the structure of
the cathode gas flow channel, so liquid water will accumulate in Nafion predicted by molecular dynamics with  = 15H2 O/SO3 . Only the water and
the catalyst layer and build up a hydrostatic pressure. The water sulfonic acid groups are shown. The voids are occupied by the TFE backbone and
pressure will increase until the water pushes through either the PFEA side chains. (B) A unit cell of a simplified model for Nafion assuming a geo-
membrane or the cathode gas diffusion layer. The pressure required metrically regular structure. The hydrophilic domains are the struts and the voids
are the hydrophobic domains. The hydraulic radius for water transport is given by a,
to push water through the Nafion membrane depends on the cur-
and d is the spacing between hydrophilic domains. (C) The connection of unit cells
rent density of the fuel cell. The hydrostatic pressure to remove to make a bicontinuous channel network.
the water by transport through the membrane is found by equat-
ing the water flux through the membrane to the water production.
EW Nafion for the hydrophilic domains to percolate through the
The hydrostatic pressure differential to force the water through the
hydrophobic matrix [19]. Above the critical threshold, proton
membrane at typical fuel cell current densities would be very large
conductivity increases quadratically with the hydrophilic volume
(equation 8).
fraction; H+ ∼(ϕ − ϕc )2 . Zhao et al. have shown that water diffu-
j/2F sion in Nafion also increases quadratically with the hydrophilic
P = ≈ 1 MPa, j = 1 A/cm2 (8)
kw /tm volume fraction; Dw ∼ ϕ2 [20]. Zhao also showed that tortuosity,
The hydraulic pressure to force water through the membrane can , of the hydrophilic domains decreases dramatically with water
be compared to the pressure to force liquid water through the GDL. volume fraction from  ∼ 20 at ϕw = 0.02 to  ∼ 2 at ϕw ∼ 0.1.
Benziger and co-workers previously measured liquid water per- The convection of water through Nafion can also serve as a probe
meation through GDL materials [43]. There is a minimum pressure to its structure. The hydrophilic domains form a network through
for liquid water penetration into the largest pores of the GDL of which water flows. The water flux depends on the size and den-
∼10 kPa. Above the penetration pressure the permeation coeffi- sity of the hydrophilic channels. Fig. 5A shows molecular dynamics
cient through the GDL is ∼10−12 cm2 , which is much greater than results from Daly for the hydrophilic phase separated domains for
kw ∼ 10−16 cm2 for permeation through the membrane. This simple 1100 EW Nafion with  = 15H2 O/SO3 [57]. The hydrophilic domains
calculation shows that most of the water formed at the cathode will form a channel network. A simplified model of the channel network
be transported through the GDL to the cathode gas flow channel. in Nafion is illustrated in Fig. 5. This model is greatly simplified,
assuming regularly spaced uniform channels. The utility of the
4.4. The hydrophilic pore network model is to make quantitative predictions of flow rates through
Nafion.
Nafion consists of hydrophilic sulfonic acid groups dispersed The microphase separated structure is approximated by a reg-
in a hydrophobic matrix of tetrafluoroethylene and perfluoroalkyl ular array of unit cells consisting of struts and cavities. The
ethers [18,24,44–53]. The sulfonic acid groups sorb water swelling struts and cavities of the unit cell represent Nafion’s phase sep-
the hydrophilic network through which water and protons are arated structure. The volume fractions of the struts and cavities
transported. The microphase separated structure of Nafion is analo- are characterized by two quantities: a = characteristic dimen-
gous to a porous medium where water may be transported through sion of the hydrophilic channel (e.g. hydrodynamic radius), and
a convoluted hydrophilic pore structure. Gierke proposed a pore d = characteristic distance between hydrophilic channels. The vol-
structure with spherical hydrophilic clusters of sulfonic acid groups ume fraction of the struts, representing water sorbed into the
and water that are connected by narrow hydrophilic channels hydrophilic channels is related to the two quantities a and d by Eq.
[44]. More refinements to the Gierke model have been proposed (9). (One could imagine the hydrophilic and hydrophobic phases
over the past 30 years based primarily on results from Small switched. Because the model is symmetric the results are indepen-
Angle X-ray scattering and Small Angle Neutron scattering [18]. dent of whether one assumed the hydrophilic phases consisted of
Those results have shown that there is a characteristic scattering the struts between vertices or the central voids.)
distance of ∼5 nm that appears when water is sorbed in Nafion  a 2  a 3
[24,27,48–51]. Most studies have associated that distance with the ϕ = 12 − 16 (9)
d d
spacing between hydrophilic clusters in the Nafion [24,48,49,54];
however, some studies have suggested that the scattering peak The quadratic term in Eq. (9) is the volume of non-intersecting
corresponds to the size of the hydrophilic clusters [55,56]. channels; the cubic term in Eq. (9) accounts for the overlapping
Transport properties can also provide structural information volumes of channels going in different directions.
about Nafion. Wu et al. have recently shown that there is a crit- When a hydraulic pressure is applied it will cause water to flow
ical hydrophilic volume fraction (ϕ) threshold of 0.1 in 1100 in the direction of the pressure gradient; the applied pressure must
Q. Duan et al. / Journal of Membrane Science 392–393 (2012) 88–94 93

Table 2
Pore network parameters for liquid water flow in Nafion.

Membrane T (◦ C) ϕw [2kw ]1/2 (nm) a (nm) d (nm)

N115 23 0.404 ± 0.03 0.27 1.12 ± 0.15 5.14 ± 0.2


N115 40 0.36 ± 0.03 0.32 1.26 ± 0.15 5.69 ± 0.2
N115 60 0.37 ± 0.03 0.33 1.24 ± 0.15 5.43 ± 0.2
N115 80 0.39 ± 0.03 0.38 1.38 ± 0.15 5.94 ± 0.2
N1110 23 0.34 ± 0.03 0.27 1.14 ± 0.15 5.25 ± 0.2
N1035 23 0.38 ± 0.03 0.34 1.28 ± 0.15 5.57 ± 0.2
N212 23 0.35 ± 0.03 0.28 1.17 ± 0.15 5.37 ± 0.2

overcome the viscous drag with the walls of the channels. The The results indicate that the spacing between hydrophilic
flow rate as a function of pressure drop may be approximated by domains, d, ranges from 5 to 6 nm, increasing with water
Poiseuille flow (Eq. (10)), where laminar flow in uniform square volume fraction. Those values agree with the characteristic dis-
channels with hydraulic radius rh is assumed. tance observed from scattering experiments (e.g. see Fig. 4).
  The hydrophilic channel diameter, dchannel = 2a, is approximately
rh2 P
Qw,pore = (2rh )2 (10) 2.3–2.7 nm.
8 Lpore The pore structure suggested here is similar to that previous
proposed by Gebel and Diat for proton transport in Nafion [48,51].
The total water flow for channels spaced at a distance d apart is
Gebel and Lambard had previously combined water sorption data
given by Eq. (11). To account for distortions of the real pore network
with the characteristic dimension from SAXS experiments to pre-
from the idealized model the effective length of the channels is
dict hydrophilic domain sizes and spacing [51]. The agreement of
given by the membrane thickness multiplied by a tortuosity factor,
the water sorption data and Porod analysis from Gebel and co-
.
   

workers was not very good assuming spherical domains. The fibril
(2rh )2 rh2 P rh4 P structure of hydrophobic cores surrounded by a hydrophobic shell
Qw = = (11)
d2 8 Lpore d2 2tm more recently proposed by Gebel and co-workers was more con-
sistent with scattering results and sorption data and we show here
Eqs. (10) and (11) assume that the Poiseuille flow description of it is also consistent with transport data.
liquid flow through cylindrical channels can be extended to more The water sorption and hydraulic water transport are not con-
complex geometries, this assumption has been validated for both sistent with models that suggest the hydrophilic pores are ∼5 nm
gas and liquid flow through porous materials, leading to the com- in diameter [56]. The flow results are also not consistent with the
monly employed Kozeny equation for flow through porous media classical Gierke model of spherical clusters connected by narrow
[58]. channels. If there were narrow channels between large spher-
The hydraulic radius accounts for the volume available for flow ical clusters the hydraulic permeability of the pores would be
divided by interfacial area over which the shear forces are exerted, greatly reduced, which would require the spacing between pores
as given by Eq. (12). The hydrophilic volume is the volume fraction to be reduced. The flow rate at fixed pressure increases with the
of water sorbed into the polymer membrane, given by Eq. (2). hydraulic radius raised to the 4th power, so a small decrease in
hydrophilic volume/unit volume ϕw channel radius would require more closely spaced hydrophilic
rh = = (12) channels.
interfacial area/unit volume ˛int
The interfacial area between the hydrophilic domains and the
hydrophobic matrix is not directly accessible for measurement.
However, the interfacial area is related to the hydrophilic network
parameters a and d, and is given by Eq. (13). Eq. (13) only includes 5. Conclusions
the interfacial area of the struts in the direction of flow.
Hydraulic permeation of water through Nafion membranes has
˛int = 24a(d − 2a) (13) been examined for different equivalent weights at temperatures
20–80 ◦ C. The principal findings are:
The hydraulic radius can also be expressed as a function of the
hydrophilic network parameters as given by Eq. (14).

3ad − 4a2
rh = (14) 1. Hydraulic permeation decreases with increasing membrane
3(d − 2a)
thickness.
Eqs. (2) and (11)–(14) can be combined to determine a and d 2. Hydraulic permeation increases with increasing temperature.
as functions of the experimentally measured and water volume 3. Hydraulic permeation increases with decreasing equivalent
fraction (ϕw ) and water flow rate (Qw ). weight of Nafion.
The water volume fractions were determined from water sorp- 4. Hydraulic permeation is of negligible importance for water per-
tion data and the results are listed in Table 2. From Eqs. (1) and vaporation; the increase of liquid water activity with pressure is
(11) the experimental permeability can be related to the hydraulic small.
radius and spacing between pores. 5. Hydraulic permeation is the dominant transport process for
water when liquid water is present on both sides of a Nafion
rh4 membrane.
= [2kw ] (15)
d2 6. Liquid water removal from the cathode catalyst layer is primarily
Assuming  = 2 from the measurements of Zhao et al. [20] values through the cathode GDL.
of a and d were determined for each membrane at the different 7. Water sorption and hydraulic permeation results suggest a
temperatures using Eqs. (9) and (15); the results are summarized hydrophilic pore network in Nafion with hydrophilic domains
in Table 2. ∼2.5 nm in diameter spaced 5.5 nm apart.
94 Q. Duan et al. / Journal of Membrane Science 392–393 (2012) 88–94

Acknowledgements [27] M. Pineri, G. Gebel, R.J. Davies, O. Diat, Water sorption–desorption in Nafion (R)
membranes at low temperature, probed by micro X-ray diffraction, Journal of
Power Sources 172 (2007) 587–596.
This work was supported by the National Science Foundation [28] M. Eikerling, Y.I. Kharkats, A.A. Kornyshev, Y.M. Volfkovich, Phenomenological
MRSEC Program through the Princeton Center for Complex Materi- theory of electro-osmotic effect and water management in polymer electrolyte
als NSF DMR-0819860. Q.D. thanks the Chinese Scholarship Council proton-conducting membranes, Journal of the Electrochemical Society 145
(1998) 2684–2699.
for fellowship support. [29] J. Divisek, M. Eikerling, V. Mazin, H. Schmitz, U. Stimming, Y.M. Volfkovich, A
study of capillary porous structure and sorption properties of Nafion proton-
exchange membranes swollen in water, Journal of The Electrochemical Society
References 145 (1998) 2677–2683.
[30] M. Eikerling, A.A. Kornyshev, U. Stimming, Electrophysical properties of poly-
[1] P.W. Majsztrik, M.B. Satterfield, A.B. Bocarsly, J.B. Benziger, Water sorption, mer electrolyte membranes: a random network model, Journal of Physical
desorption and transport in Nafion membranes, Journal of Membrane Science Chemistry B 101 (1997) 10807–10820.
301 (2007) 93–106. [31] M.H. Eikerling, P. Berg, Poroelectroelastic theory of water sorption and swelling
[2] B. Kientiz, H. Yamada, N. Nonoyama, A.Z. Weber, Interfacial water transport in polymer electrolyte membranes, Soft Matter 7 (2011) 5976–5990.
effects in proton-exchange membranes, Journal of Fuel Cell Science and Tech- [32] P. Majsztrik, A. Bocarsly, J. Benziger, Water permeation through nafion mem-
nology 8 (2011). branes: the role of water activity, Journal of Physical Chemistry B 112 (2008)
[3] S.H. Ge, X.G. Li, B.L. Yi, I.M. Hsing, Absorption, desorption, and transport of water 16280–16289.
in polymer electrolyte membranes for fuel cells, Journal of The Electrochemical [33] M. Adachi, T. Navessin, Z. Xie, S. Holdcroft, POLY 364-Ex-situ water permeation
Society 152 (2005) A1149–A1157. measurements through proton exchange membranes, Abstracts of Papers of
[4] M. Adachi, T. Navessin, Z. Xie, B. Frisken, S. Holdcroft, Correlation of in situ the American Chemical Society 238 (2009).
and ex situ measurements of water permeation through Nafion NRE211 pro- [34] P.W. Majsztrik, A.B. Bocarsly, J.B. Benziger, Viscoelastic response of nafion.
ton exchange membranes, Journal of the Electrochemical Society 156 (2009) effects of temperature and hydration on tensile creep, Macromolecules 41
B782–B790. (2008) 9849–9862.
[5] M. Adachi, T. Navessin, Z. Xie, F.H. Li, S. Tanaka, S. Holdcroft, Thickness depen- [35] M.B. Satterfield, J.B. Benziger, Viscoelastic properties of nafion at elevated tem-
dence of water permeation through proton exchange membranes, Journal of perature and humidity, Journal of Polymer Science Part B: Polymer Physics 47
Membrane Science 364 (2010) 183–193. (2009) 11–24.
[6] V. Freger, E. Korin, J. Wisniak, E. Korngold, M. Ise, K.D. Kreuer, Diffusion of water [36] G. Gebel, O. Diat, S. Escribano, R. Mosdale, Water profile determination in a
and ethanol in ion-exchange membranes: limits of the geometric approach, running PEMFC by small-angle neutron scattering, Journal of Power Sources
Journal of Membrane Science 160 (1999) 213–224. 179 (2008) 132–139.
[7] J.T. Hinatsu, M. Mizuhata, H. Takenaka, Water-uptake of perfluorosulfonic acid [37] G. Gebel, S. Lyonnard, H. Mendil-Jakani, A. Morin, The kinetics of water sorp-
membranes from liquid water and water-vapor, Journal of The Electrochemical tion in Nafion membranes: a small-angle neutron scattering study, Journal of
Society 141 (1994) 1493–1498. Physics-Condensed Matter 23 (2011).
[8] C.W. Monroe, T. Romero, W. Merida, M. Eikerling, A vaporization-exchange [38] Z. Zhang, A.E. Marble, B. MacMillan, K. Promislow, J. Martin, H. Wang, B.J.
model for water sorption and flux in Nafion, Journal of Membrane Science 324 Balcom, Spatial and temporal mapping of water content across Nafion mem-
(2008) 1–6. branes under wetting and drying conditions, Journal of Magnetic Resonance
[9] S. Motupally, A.J. Becker, J.W. Weidner, Water transport in polymer electrolyte 194 (2008) 245–253.
membrane electrolyzers used to recycle anhydrous HCl. I. Characterization of [39] Z. Zhang, K. Promislow, J. Martin, H. Wang, B.J. Balcom, Bi-modal water trans-
diffusion and electro-osmotic drag, Journal of The Electrochemical Society 149 port behavior across a simple Nafion membrane, Journal of Power Sources 196
(2002) D63–D71. (2011) 8525–8530.
[10] M.B. Satterfield, J.B. Benziger, Non-fickian water vapor sorption dynamics by [40] S.K. Park, E.A. Cho, I.H. Oh, Characteristics of membrane humidifiers for polymer
Nafion membranes, Journal of Physical Chemistry B 112 (2008) 3693–3704. electrolyte membrane fuel cells, Korean Journal of Chemical Engineering 22
[11] T. Takamatsu, M. Hashiyama, A. Eisenberg, Sorption phenomena in Nafion (2005) 877–881.
membranes, Journal of Applied Polymer Science 24 (1979) 2199–2220. [41] Perma Pure LLC, http://www.permapure.com/industry/enviro/technology/,
[12] S. Tsushima, K. Teranishi, S. Hirai, Water diffusion measurement in fuel-cell SPE 2011.
membrane by NMR, Energy 30 (2005) 235–245. [42] Cell Craft, http://www.cellkraft.se/humidity and steam/home.html, 2011.
[13] A.Z. Weber, J. Newman, Transport in polymer-electrolyte membranes. I. Phys- [43] J. Benziger, J. Nehlsen, D. Blackwell, T. Brennan, J. Itescu, Water flow in the
ical model, Journal of The Electrochemical Society 150 (2003) A1008–A1015. gas diffusion layer of PEM fuel cells, Journal of Membrane Science 261 (2005)
[14] Q.G. Yan, H. Toghiani, J.X. Wu, Investigation of water transport through mem- 98–106.
brane in a PEM fuel cell by water balance experiments, Journal of Power Sources [44] T.D. Gierke, W.Y. Hsu, The cluster-network model of ion clustering in perfluo-
158 (2006) 316–325. rosulfonated membranes, ACS Symposium Series 180 (1982) 283–307.
[15] T.A. Zawodzinski, C. Derouin, S. Radzinski, R.J. Sherman, V.T. Smith, T.E. [45] T.D. Gierke, G.E. Munn, F.C. Wilson, The morphology in Nafion perfluori-
Springer, S. Gottesfeld, Water-uptake by and transport through Nafion(R) 117 nated membrane products, as determined by wide-angle and small-angle
membranes, Journal of The Electrochemical Society 140 (1993) 1041–1047. X-ray studies, Journal of Polymer Science Part B: Polymer Physics 19 (1981)
[16] T.A. Zawodzinski, M. Neeman, L.O. Sillerud, S. Gottesfeld, Determination of 1687–1704.
water diffusion-coefficients in perfluorosulfonate ionomeric membranes, Jour- [46] T.D. Gierke, G.E. Munn, F.C. Wilson, Morphology of perfluorosulfonated mem-
nal of Physical Chemistry 95 (1991) 6040–6044. brane products – wide-angle and small-angle X-ray studies, ACS Symposium
[17] M.W. Verbrugge, E.W. Schneider, R.S. Conell, R.F. Hill, The effect of Series 180 (1982) 195–216.
temperature on the equilibrium and transport-properties of saturated [47] A. Eisenberg, Clustering of ions in organic polymers. A theoretical approach,
poly(perfluorosulfonic acid) membranes, Journal of The Electrochemical Soci- Macromolecules 3 (1970) 147–154.
ety 139 (1992) 3421–3428. [48] O. Diat, G. Gebel, Proton channels, Nature Materials 7 (2008) 13–14.
[18] K.A. Mauritz, R.B. Moore, State of understanding of Nafion, Chemical Reviews [49] G. Gebel, P. Aldebert, M. Pineri, Swelling study of perfluorosulphonated
104 (2004) 4535–4585. ionomer membranes, Polymer 34 (1993) 333–339.
[19] X. Wu, X. Wang, G. He, J. Benziger, Differences in water sorption and proton [50] G. Gebel, Structural evolution of water swollen perfluorosulfonated ionomers
conductivity between Nafion and SPEEK, Journal of Polymer Science Part B: from dry membrane to solution, Polymer 41 (2000) 5829–5838.
Polymer Physics 49 (2011) 1437–1445. [51] G. Gebel, J. Lambard, Small-angle scattering study of water-swollen perfluori-
[20] Q. Zhao, P. Majsztrik, J. Benziger, Diffusion and interfacial transport of water in nated ionomer membranes, Macromolecules 30 (1997) 7914–7920.
Nafion, Journal of Physical Chemistry B 115 (2011) 2717–2727. [52] K.D. Kreuer, On the development of proton conducting polymer membranes
[21] J. Peron, A. Mani, X.S. Zhao, D. Edwards, M. Adachi, T. Soboleva, Z.Q. Shi, Z. Xie, T. for hydrogen and methanol fuel cells, Journal of Membrane Science 185 (2001)
Navessin, S. Holdcroft, Properties of Nafion (R) NR-211 membranes for PEMFCs, 29–39.
Journal of Membrane Science 356 (2010) 44–51. [53] M. Pineri, A. Eisenberg, Structure and properties of ionomers, in: NATO ASI
[22] R.A. Register, T.R. Bell, Miscible blends of zinc-neutralized sulfonated Series. Series C: Mathematical and Physical Sciences, Reidel Pub. Co., Norwell,
polystyrene and poly(2,6-dimethyl 1,4-phenylene oxide), Journal of Polymer MA, 1987.
Science Part B: Polymer Physics 30 (1992) 569–575. [54] K. Schmidt-Rohr, Q. Chen, Parallel cylindrical water nanochannels in Nafion
[23] C. Yang, S. Srinivasan, A.B. Bocarsly, S. Tulyani, J.B. Benziger, A comparison of fuel-cell membranes, Nature Materials 7 (2008) 75–83.
physical properties and fuel cell performance of Nafion and zirconium phos- [55] M. Fujimura, T. Hashimoto, H. Kawai, Small-angle X-ray-scattering study of
phate/Nafion composite membranes, Journal of Membrane Science 237 (2004) perfluorinated ionomer membranes. 1. Origin of 2 scattering maxima, Macro-
145–161. molecules 14 (1981) 1309–1315.
[24] G. Gebel, O. Diat, Neutron and X-ray scattering: suitable tools for studying [56] M. Fujimura, T. Hashimoto, H. Kawai, Small-angle X-ray-scattering study of
ionomer membranes, Fuel Cells 5 (2005) 261–276. perfluorinated ionomer membranes. 2. Models for ionic scattering maximum,
[25] K.A. Page, F.A. Landis, A.K. Phillips, R.B. Moore, SAXS analysis of the thermal Macromolecules 15 (1982) 136–144.
relaxation of anisotropic morphologies in oriented Nafion membranes, Macro- [57] K. Daly, Structure and Properties of Nafion from Molecular Simulations, Prince-
molecules 39 (2006) 3939–3946. ton University, Princeton, NJ, 2011.
[26] L. Rubatat, O. Diat, Stretching effect on Nafion fibrillar nanostructure, Macro- [58] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, 2nd ed., Wiley,
molecules 40 (2007) 9455–9462. New York, 2002.

Potrebbero piacerti anche